Anda di halaman 1dari 12

Z. Cai, L.-J. Yan / Journal of Biochemical and Pharmacological Research, Vol.

1 (1): 15-26, March 2013

Review

Protein Oxidative Modifications: Beneficial


Roles in Disease and Health
Zhiyou Caia, Liang-Jun Yanb,*

Abstract: Protein oxidative modifications, also known as


protein oxidation, are a major class of protein
posttranslational modifications. They are caused by
reactions between protein amino acid residues and reactive
oxygen species (ROS) or reactive nitrogen species (RNS)
and can be classified into two categories: irreversible
modifications and reversible modifications. Protein
oxidation has been often associated with functional decline
of the target proteins, which are thought to contribute to
normal aging and age-related pathogenesis. However, it has
now been recognized that protein oxidative modifications
can also play beneficial roles in disease and health. This
review summarizes and highlights certain positive roles of
protein oxidative modifications that have been documented
in the literature. Covered oxidatively modified protein
adducts
include
carbonylation,
3-nitrotyrosine,
s-sulfenation, s-nitrosylation, s-glutathionylation, and
disulfide formation. All of which have been widely analyzed
in numerous experimental systems associated with redox
stress conditions. The authors believe that selected protein
targets, when modified in a reversible manner in
prophylactic approaches such as preconditioning or
ischemic tolerance, may provide potential promise in
maintaining health and fighting disease.
Keywords: carbonylation, cysteine, glutathionylation,
ischemic tolerance, nitrosylation, nitrotyrosine, sulfenation,
sulfenic acid, postconditioning, preconditioning, oxidative
modifications, reactive oxygen species

__________________________________________
a

Department of Neurology, Lu'an People's Hospital, the Lu'an


Affiliated Hospital of Anhui Medical University, Lu'an, Anhui
Province, China, 237005
b
Department of Pharmacology and Neuroscience and Institute for
Aging and Alzheimer's Disease Research, University of North Texas
Health Science Center, Fort Worth, Texas, USA
* Corresponding author. Tel: +1 817-735-2386; Fax: +1 817-735-0408
Email: liang-jun.yan@unthsc.edu
(Received January 10, 2013; Revised February 4, 2013; Accepted
February 4, 2013; Published online February 5, 2013)

ISSN 2168-8761 print/ISSN 2168-877X online

1. Introduction

n order to cope with environmental challenges, cells rely on a


variety of posttranslational modification mechanisms to
expand protein function [1-4]. Of all the documented
posttranslational modifications, oxidative modification of the
side chains of various amino acid residues forms a major
category of protein posttranslational modifications [5-7].
Protein oxidative modifications can be generally classified into
two categories: irreversible oxidation and reversible oxidation
[8-10]; both of which can be selectively induced by reactive
oxygen species (ROS) and reactive nitrogen species (RNS) [11,
12].
Earlier studies of protein oxidation nearly exclusively
focused on the detrimental effects of protein oxidation in aging
and diseases [5, 9, 13-17]. It has now been firmly recognized,
however, that protein oxidation can also play a positive role in
many cellular functions. This gradual realization of the
beneficial roles of protein oxidation may be attributed to
accumulating evidence that ROS and RNS are indispensible for
cell survival [18-22] and regeneration [23], and in many cases,
they are required for recovery of cellular functions by creating
positive stress conditions whereby cell survival mechanisms
are reprogrammed to extend life span [24-27] or to withstand
severe, or lethal challenges [28-31].
In this article, we review both irreversible and reversible
oxidative modifications that are beneficial in health and
disease. Modification adducts to be discussed include protein
carbonyls, 3-nitrotyrosine, and cysteine oxidation products
(Fig. 1). As protein cysteine residue is the one that often
undergoes reversible redox modifications by ROS or RNS
[32-34], we have inclined to devote more space on cysteine
modifications including s-sulfenation, s-nitrosylation,
s-glutathionylation, and disulfide formation that are all
reversible [35-38]. It should be noted that protein oxidative
modifications that have deleterious effects in health and disease
are beyond the scope of this review and will only be
sporadically mentioned.

~ 15 ~

http://www.researchpub.org/journal/jbpr/jbpr.html

Z. Cai, L.-J. Yan / Journal of Biochemical and Pharmacological Research, Vol. 1 (1): 15-26, March 2013

2. Cellular sources of oxidants


There are many systems inside a cell that can generate ROS.
Mitochondria are recognized as the major site for ROS
production [39-41]; and both complexes I and III have been
established to be the specific sites for mitochondrial ROS
generation [42-45]. Besides mitochondria, many enzymes are
also capable of producing ROS. These include, but not limited
to, NADPH oxidase [46, 47], xanthine oxidase [48, 49],
-ketoglutarate dehydrogenase complex [50-52], d-amino acid
oxidases [53-55], and dihydrolipoamide dehydrogenase
[56-62]. On the other hand, nitric oxide production in vivo is
mainly achieved by nitric oxide synthases [63-65] though under
certain conditions deoxygenated myoglobin [66] or xanthine
oxidoreductase [67] or cytochrome c oxidase [68] can be
involved in NO production; and in vitro nitric oxide donors are
also frequently used either in experimental systems [69-71] or
for therapeutic purpose [72-74]. It should be noted that in the
presence of superoxide anion, nitric oxide can rapidly react
with superoxide anion to yield peroxynitrite [75-77], a reactive
species that is highly reactive toward redox-sensitive amino
acid residues including tyrosine and cysteine [78, 79].

Nitrotyrosine, usually 3-nitrotyrosine, is formed between


reactive nitrogen species and a proteins tyrosine residue [78,
98, 99]. This modification is a highly selective process as not all
proteins or all tyrosine residues on a target protein can get
nitrated [100]. Formation of nitrotyrosine is often thought to be
accompanied with acute or chronic inflammation disease
[101-104], whereby level of nitric oxide is elevated [102,
104-106]. While numerous studies have investigated the
deleterious effects of 3-nitrotyrosines [107, 108], concurrent
with development of methods for detection and quantitation
[109, 110], this modification has been detected under normal
physiological conditions such as healthy pregnancy [111, 112],
indicating that formation of 3-nitrotyrosine has physiological
function.
Carbonyls (carbonylation)

3-nitrotyrosine

Irreversible

Protein
Reversible

3. Irreversible protein oxidative modifications


First, we would like to discuss briefly the possible beneficial
role of irreversible modifications. These types of modifications
include mainly protein carbonylation and tyrosine nitration [11,
80-84]. Both modifications are often associated with oxidative
damage and have been used as biomarkers for assessment of
oxidative stress in aging and diseases [13, 15-17, 85]. While
both carbonylation and nitration can have detrimental effects
on the target proteins, evidence has also emerged that such
modifications can also play positive roles in cellular function
under stress conditions.
3.1. Protein carbonyls
Protein carbonyls formed on several amino acids residues,
including arginine, histidine, lysine, proline, threonine and
cysteine, are the most widely used biomarker for measurement
of protein oxidation and oxidative stress in aging and diseases
[5, 8, 11-14, 86-90]. As the modification occurs on multiple
amino acid residues on selected protein targets [15-17, 91], its
magnitude is much greater than any other modifications that
occur only on a specific amino acid residue [11, 12], and thus is
more readily detectable. Many studies have employed protein
carbonylation to evaluate the detrimental effects of protein
oxidation and oxidative stress [13-16, 87-90]; evidence for
positive effects of this modification, however, has started to
accumulate. For example, protein carbonylation has been
shown to be involved in signal transduction [92-95] and is
known to be involved in ischemic preconditioning eliciting
protection against reperfusion-induced tissue injuries [96, 97].
3.2. Protein nitrotyrosine

ISSN 2168-8761 print/ISSN 2168-877X online

Sulfenic acid
(s-sulfenation)

disulfides
s-nitrosothiols
S-nitrosylation)
s-glutathione
(s-glutathionylation)

Fig. 1. Irreversible and reversible protein oxidation products discussed


in this review. Irreversible oxidation includes protein carbonyls and
3-nitrotyrosine while reversible oxidation includes cysteine
modification products such as sulfenic acid, nitrosothiols, and
s-glutathione.

4. Reversible protein oxidative modifications: protein


cysteine modifications
4.1. Chemistry of protein cysteine residues
At neutral pH under physiological conditions, free cysteine
residues have a pKa value that is around 8.5, which makes
oxidative modifications impossible [113]. To be susceptible to
oxidation, the pKa value of a cysteine residue needs to be lower
than the physiological pH value (pH 7.4), a condition under
which, the cysteine SH group becomes thiolated (thiolate
anion) [113-115]. It is those thiolated cysteine residues that are
redox reactive [35, 116]. This thiolation process, decreasing the
pKa value to 7.2 or lower, can be achieved via many factors
such as hydrogen bonding [117, 118], the effect of adjacent
basic amino acid residues [117], the microenvironment of the
target cysteine residues [117], and substrate binding [119]. For

~ 16 ~

http://www.researchpub.org/journal/jbpr/jbpr.html

Z. Cai, L.-J. Yan / Journal of Biochemical and Pharmacological Research, Vol. 1 (1): 15-26, March 2013

example, albumin cysteine 34 has a very low pKa value of 5


[120]. Hence under physiological condition, it exists as thiolate
anion and is very reactive towards oxidants, thiols, metals, and
disulfides [121-123].
As described above, thiols with low pKa values are more
reactive because they are usually deprotonated or thiolated at
physiological pH [124-126]. Therefore, oxidation of protein
cysteines that are redox reactive is also a highly selective
process [127, 128]. As shown in Fig. 2, cysteine oxidation
usually starts with the formation of sulfenic acid, from which a
variety of oxidation products can be furtherly formed and many
of them are reversible and well defined chemically. These
cysteine oxidation products include disulfide formation (S-S-),
S-glutathionylation (protein-SSG), S-nitrosylation (-SNO),
sulfenic acid formation (-SOH, or S-sulfenation) and have all
been demonstrated in redox regulation of protein functions by
ROS and RNS [35, 129, 130]. Importantly, all of which have
been implicated to play beneficial roles in disease and health
because they may protect the target proteins from further
oxidation that will otherwise permanently damage the target
proteins [131-133]. Another mechanism is that these
modifications also play a role in redox signaling cascades that
boost cellular defense systems to better counteract stress insults
[134-136].

peroxynitrite [38, 129, 137, 142, 143]. Although being a simple


chemical modification, sulfenic acid formation can have a
dramatic effect on protein function [130, 137, 144]. It was for a
long time regarded as an intermediate, unstable cysteine
oxidation product, which may still be true for many proteins
[137, 145, 146]. Growing evidence, however, has demonstrated
that stable-SOH indeed exists, making trapping, labeling,
detecting, and quantitating possible for further evaluation of the
formed SOH [142, 147-150]. A beneficial effect of protein
SOH formation has been elegantly demonstrated in studies
whereby s-sulfenation of aldose reductase protects the heart
against ischemic/reperfusion injury [151-153]. Specifically,
these studies found that cyse-298s sulfenation of aldose
reductase by peroxynitrite shows great protection against
cardiac ischemic injury; and administration of peroxynitrite
scavengers not only eliminates cys-298s sulfenation, but also
abolishes cardiac protection against ischemic injury. In
unrelated studies, Michalek et al. demonstrated that protein
sulfenation is indispensible for T-cell growth and proliferation
as arrest of sulfenic acids greatly impairs T cell maturation
[154]. Another example of a beneficial role of P-SOH is that of
the sulfenation of nitrile hydratase; sulfenic acid formation on
this enzymes Cys114 residue is absolutely essential for the
enzymes catalytic activity [155].
4.3. Protein s-nitrosylation

S-S-G
SNO

Reversible

GSH

H2 O
HNO

NO

HS

.
S

ROS

S-S

SOH

Irreversible

ROS

SO2H

ROS

SO3H

Fig. 2. Chemistry of cysteine oxidative modifications. Sulfenic acid is


truly an intermediate product during cysteine oxidation. Given
appropriate conditions, s-nitrosothiols can also be discomposed to
yield sulfenic acids with concurrent production of nitroxyl [137].
Sulfenic acid can be further oxidized to form disulfide bonds,
s-glutathionylation. Irreversible oxidation products sulfinic and
sulfonic acids are also shown.

Protein s-nitrosylation can be induced by nitric oxide,


nitroxyl, and peroxynitrite [156, 157]. This modification has
been regarded as functionally equivalent to protein
phosphorylation and dephosphorylation [158-160]. Besides
occurring on cysteine residues other than on tyrosine, serine, or
threonine residues, s- nitrosylation is also potentially different
from phosphorylation in that nitrosylation may not involve a
delicate network consisting of kinases or enzymes that catalyze,
respectively, nitrosylation and denitrosylation, though the
existence of denitrosylases, including Cu,Zn-superoxide
dismutase and bilirubin, has been reported [161-165].
Nonetheless, s-nitrosylation has been demonstrated to be a key
modification of cysteine residues under a variety of
physiological and pathophysiological conditions [157, 166,
167]. In particular, in connection with nitric oxide-based redox
regulation of protein function, s-nitrosylation has been found to
be involved in protective mechanisms in many disorders [157,
168-170]. For example, Sheng et al. have demonstrated that
chemically-enhanced s-nitrosylation can improve recovery
from subarachnoid hemorrhage [171], and Penna et al. have
demonstrated that protein s-nitrosylation is favorably produced
during cardiac postconditioning [172].
4.4. Protein s-glutathionylation

4.2. Protein sulfenic acid formation (S-sulfenation)


This sulfur-hydroxylation product (P-SOH) possesses
powerful redox chemistry and has been demonstrated to play a
key redox regulatory role in a growing number of proteins [34,
138-141]. Its formation is mainly induced by ROS such as
hydrogen peroxide, alkyl hydroperoxides, and RNS such as
ISSN 2168-8761 print/ISSN 2168-877X online

Protein
cysteine
residues
can
also
undergo
s-glutathionylation under oxidative stress conditions [173-175].
Glutathione (GSH) is the major cellular antioxidant, yet, it can
also modify proteins via mixed disulfide formation (P-S-S-G),
leading to functional changes of the target proteins [176]. This
reversible oxidation of critical cysteine residues on proteins has

~ 17 ~

http://www.researchpub.org/journal/jbpr/jbpr.html

Z. Cai, L.-J. Yan / Journal of Biochemical and Pharmacological Research, Vol. 1 (1): 15-26, March 2013

been found to be involved in oxidative signal transduction,


control of gene expression, cell proliferation, apoptosis, and
cellular responses to protecting key regulatory molecules from
oxidative insults [173, 176-178]. Similar to s-sulfenation and
s-nitrosylation, protein-S-S-G is also often associated with a
detrimental effect on the target proteins function [179-182],
but can also protect the target protein from irreversible and
permanent
damage
[183-186].
Therefore,
protein
glutathionylation has increasingly gained great attention as a
possible means of redox regulation of protein functions in
response to oxidative stress under physiological and
pathophysiological conditions [185, 187]. For example, actin
glutathionylation
regulates
actin
dynamics
in
polymorphonuclear neutrophils [188], manipulation of
uncoupling protein 2s glutathionylation may provide a strategy
for cancer treatment [189], and glutathionylation of adenine
nuclear translocase induced by preconditioning can prevent
mitochondrial membrane permeabilization and apoptosis [190].
4.5. Protein disulfides
This is different from protein s-glutathionylation, where a
mixed disulfide between GSH and a protein-linked cysteine
residue is formed [191-193]. Native disulfide bond formation is
usually involved in correct protein folding and is catalyzed by
disulfide isomerase in the endoplasmic reticulum and the
mitochondrial intermembrane space [194-197], and should be
considered different from those formed under oxidative stress
or pathophysiological conditions. Hence herein, disulfide
formation is strictly meant to reflect inter- or intra- protein
disulfide formation that is caused by ROS or RNS [198-205].
Disulfide bonds formed between free cysteine residues upon
oxidative stress have been reported to play a beneficial role in
cellular defense systems against a variety of stress challenges
[191, 206-210]. For example, intra-protein disulfide formation
in Cdc25c upon hydrogen peroxide exposure regulates the
stability of the protein [211], and in the brain type creatine
kinase, disulfide formation between two cysteine residues
(cys74 and cys254) can serve as a cellular defense mechanism
[212]. Additionally and importantly, it is well established that
formation of disulfide linkage within Keap1 in response to
cellular stimuli by electrophiles and oxidants [213-217] is
essential for activation of the NF-E2-related factor 2 (Nrf2) that
then upregulates the expression of phase II antioxidant
enzymes under a variety of physiological and
pathophysiological conditions [218-224].
5. Protein oxidative modifications and ischemic tolerance
Posttranslational protein oxidative modifications, in
particular cysteine modifications, have been implicated in
ischemic tolerance or preconditioning [168, 225-231]. Ischemic
tolerance constitutes a positive stress that reporgrams cellular
defense systems to prevent subsequent lethal injuries
[232-237]. The phenomenon of preconditioning seems to be
universal as all tissues in mamalian systems as well as all
organisms can be preconditioned. In particular, the heart and
the brain can be preconditioned by a variety of mechanisms to
ISSN 2168-8761 print/ISSN 2168-877X online

prevent further injuries caused by ischemia reperfusion [232,


238]. Therefore, preconditioning has both prophylactic and
therapeutic value. Despite intensive studies, the mechanisms of
preconditioning has not been well understood. Nonetheless,
ROS are known to be the key molecules involved in
preconditioning development [239-243] as antioxidants
administered during induction of preconditioning can block the
preconditioning
effect
[28,
30].
Moreover,
a
moderately-elevated level of ROS, in particular, H2O2, has been
shown to be neuroprotective [221, 244-246]. Nevertheless, how
ROS work in preconditioning induction and tissue protection
remains elusive. As ROS can impart their effects by modifying
proteins, identification of endogenous protein targets of ROS
may elucidate mechanisms of protection induced by ischemic
tolerance. It is thus conceivable that identification of
oxidatively modified protein targets, especially those that can
undergo reversible oxidative modifications, may provide
insights into novel therapeutic strategies for ischemic tolerance.
It is also worth mentioning that a concept of postconditioning,
whereby the reperfusion procedure can be disrupted and
intervened to elicit protection against lethal injury, has been
recently established [247-250]. We think that postconditioning
can also be placed under the notion of ischemic tolerance. In
fact, preconditioning and postconditioning may share similar
pathways or mechanisms [250-254].
So why could protein oxidation, in particular, reversible
oxidation-induced by ischemic tolerance be involved in
protection against subsequent ischemic injury? As it is the
reperfusion stage that often incurs the injury due to a sudden
burst in ROS production concurrent with resupply of oxygen
[255-259], oxidized proteins with altered protein function
could slow down the rate of ROS production during reperfusion
and hence could attenuate ischemic injury [230, 260]. In
addition, as already pointed out earlier in this review, oxidized
proteins induced by ischemic tolerance could also be involved
in eliciting cellular defense systems to protect against further
severe ischemia reperfusion injury [261, 262].
6. Summary and Perspectives
While studies on the detrimental or deleterious effects of
protein oxidative modifications are, and will still be,
dominating the field of protein oxidation, investigation of the
beneficial roles of protein oxidation appears to be gaining
increasing interest [135, 263]. For beneficial purposes, efforts
should be focused on proteomic identification of reversibly
oxidized proteins that may exhibit protective effects. Further,
studies on a comprehensive understanding of the mechanisms
or pathways that regulate the reversible nature of the
corresponding modifications should be undertaken. This should
be particularly true for reversible cysteine oxidation, which not
only reflects changes in cellular redox state, but can also protect
the target proteins from further damage. Additionally,
reversible cysteine oxidation is also involved in redox signaling
cascades [264-267] that can elicit positive stress responses to
prevent unpredicted disastrous events such as stroke and heart
attack. Therefore, equal efforts will also be needed to identify
those protein targets that undergo reversible cysteine

~ 18 ~

http://www.researchpub.org/journal/jbpr/jbpr.html

Z. Cai, L.-J. Yan / Journal of Biochemical and Pharmacological Research, Vol. 1 (1): 15-26, March 2013

modifications in preventative or protective approaches as such


identified protein targets may provide therapeutic values in
fighting diseases, in particular, ischemia associated cerebral
and cardiovascular diseases.
Acknowledgments

[16]
[17]

[18]

The authors wish to apologize to those whose work could


not be cited due to space limitations. LJY was supported in part
by the National Institutes of Health (Grant: AG022550) and by
the University of North Texas Health Science Center
(UNTHSC-UAEM seed grant: RI6044).

[19]

[20]

Conflict of interest: None declared.


[21]

References
[1]
[2]
[3]

[4]
[5]
[6]
[7]
[8]
[9]

[10]

[11]
[12]
[13]
[14]
[15]

Glazer AN, Delange RJ and Sigman DS. Chemical


Modification of proteins: Selected methods and analytical
procedures. Elsevier Biomedical Press, Amsterdam, 1975.
Johnson BC. Posttranslational covalent modifications of
proteins. Academic Press, New York, 1983.
Graves DJ, Martin BL and Wang JH. Co- and
post-translational modification of proteins: Chemical
principles and biological effects. Oxford University Press,
Oxford, 1994.
Walsh CT. Posttranslational modification of proteins:
Expanding nature's inventory. Robert and Company
Publishers, Eaglewood, Colorado, 2006.
Stadtman ER. Protein oxidation and aging. Science 1992;
257:1220-1224.
Shacter E. Quantification and significance of protein oxidation
in biological samples. Drug Metab Rev 2000; 32:307-326.
Shacter E. Protein oxidative damage. Methods Enzymol 2000;
319:428-436.
Stadtman ER and Levine RL. Free radical-mediated oxidation
of free amino acids and amino acid residues in proteins. Amino
Acids 2003; 25:207-218.
Levine RL, Garland D, Oliver CN, Amici A, Climent I, Lenz
AG, Ahn BW, Shaltiel S and Stadtman ER. Determination of
carbonyl content in oxidatively modified proteins. Methods
Enzymol 1990; 186:464-478.
Fedorova M, Kuleva N and Hoffmann R. Reversible and
irreversible modifications of skeletal muscle proteins in a rat
model of acute oxidative stress. Biochim Biophys Acta 2009;
1792:1185-1193.
Yan LJ. Analysis of oxidative modification of proteins. Curr
Protoc Protein Sci 2009; Chapter 14:Unit14 4.
Yan LJ and Forster MJ. Chemical probes for analysis of
carbonylated proteins: A review. J Chromatogr B Analyt
Technol Biomed Life Sci 2011; 879:1308-1315.
Stadtman ER. Protein oxidation in aging and age-related
diseases. Ann N Y Acad Sci 2001; 928:22-38.
Stadtman ER. Protein oxidation and aging. Free Radic Res
2006; 40:1250-1258.
Yan LJ, Levine RL and Sohal RS. Oxidative damage during
aging targets mitochondrial aconitase. Proc. Natl. Acad. Sci.
USA 1997; 94:11168-11172.

ISSN 2168-8761 print/ISSN 2168-877X online

[22]
[23]

[24]

[25]
[26]

[27]

[28]

[29]
[30]

[31]

[32]

~ 19 ~

Yan LJ and Sohal RS. Mitochondrial adenine nucleotide


translocase is modified oxidatively during aging. Proc Natl
Acad Sci USA 1998; 95:12896-12901.
Yan LJ, Christians ES, Liu L, Xiao X, Sohal RS and Benjamin
IJ. Mouse heat shock transcription factor 1 deficiency alters
cardiac redox homeostasis and increases mitochondrial
oxidative damage. EMBO J 2002; 21:5164-5172.
Groeger G, Quiney C and Cotter TG. Hydrogen peroxide as a
cell-survival signaling molecule. Antioxid Redox Signal 2009;
11:2655-2671.
Knoefler D, Thamsen M, Koniczek M, Niemuth NJ, Diederich
AK and Jakob U. Quantitative in vivo redox sensors uncover
oxidative stress as an early event in life. Mol Cell 2012;
47:767-776.
Groeger G, Doonan F, Cotter TG and Donovan M. Reactive
oxygen species regulate prosurvival ERK1/2 signaling and
bFGF expression in gliosis within the retina. Invest
Ophthalmol Vis Sci 2012; 53:6645-6654.
Bevilacqua E, Gomes SZ, Lorenzon AR, Hoshida MS and
Amarante-Paffaro AM. NADPH oxidase as an important
source of reactive oxygen species at the mouse maternal-fetal
interface: putative biological roles. Reprod Biomed Online
2012; 25:31-43.
Watson J. Oxidants, antioxidants and the current incurability of
metastatic cancers. Open Biol 2013; 3:120144.
Love NR, Chen Y, Ishibashi S, Kritsiligkou P, Lea R, Koh Y,
Gallop JL, Dorey K and Amaya E. Amputation-induced
reactive oxygen species are required for successful Xenopus
tadpole tail regeneration. Nat Cell Biol 2013; 15:222-229.
Ristow M and Zarse K. How increased oxidative stress
promotes longevity and metabolic health: The concept of
mitochondrial hormesis (mitohormesis). Exp Gerontol 2010;
45:410-418.
Ristow M and Schmeisser S. Extending life span by increasing
oxidative stress. Free Radic Biol Med 2011; 51:327-336.
Kaneko YS, Ota A, Nakashima A, Mori K, Nagatsu I and
Nagatsu T. Regulation of oxidative stress in long-lived
lipopolysaccharide-activated microglia. Clin Exp Pharmacol
Physiol 2012; 39:599-607.
Mesquita A, Weinberger M, Silva A, Sampaio-Marques B,
Almeida B, Leao C, Costa V, Rodrigues F, Burhans WC and
Ludovico P. Caloric restriction or catalase inactivation extends
yeast chronological lifespan by inducing H2O2 and superoxide
dismutase activity. Proc Natl Acad Sci U S A 2010;
107:15123-15128.
Mori T, Muramatsu H, Matsui T, McKee A and Asano T.
Possible role of the superoxide anion in the development of
neuronal tolerance following ischaemic preconditioning in
rats. Neuropathol Appl Neurobiol 2000; 26:31-40.
Das DK, Maulik N, Sato M and Ray PS. Reactive oxygen
species function as second messenger during ischemic
preconditioning of heart. Mol Cell Biochem 1999; 196:59-67.
Puisieux F, Deplanque D, Bulckaen H, Maboudou P, Gele P,
Lhermitte M, Lebuffe G and Bordet R. Brain ischemic
preconditioning is abolished by antioxidant drugs but does not
up-regulate superoxide dismutase and glutathion peroxidase.
Brain Res 2004; 1027:30-37.
Bhagatte Y, Lodwick D and Storey N. Mitochondrial ROS
production and subsequent ERK phosphorylation are
necessary for temperature preconditioning of isolated
ventricular myocytes. Cell Death Dis 2012; 3:e345.
Beinert H. A tribute to sulfur. Eur J Biochem 2000;
267:5657-5664.

http://www.researchpub.org/journal/jbpr/jbpr.html

Z. Cai, L.-J. Yan / Journal of Biochemical and Pharmacological Research, Vol. 1 (1): 15-26, March 2013

[33]
[34]
[35]

[36]
[37]

[38]
[39]
[40]
[41]
[42]
[43]

[44]
[45]
[46]
[47]

[48]
[49]

[50]

[51]

[52]

Leonard SE and Carroll KS. Chemical 'omics' approaches for


understanding protein cysteine oxidation in biology. Curr Opin
Chem Biol 2011; 15:88-102.
Thamsen M and Jakob U. The redoxome: Proteomic analysis
of cellular redox networks. Curr Opin Chem Biol 2011;
15:113-119.
Janssen-Heininger YM, Mossman BT, Heintz NH, Forman HJ,
Kalyanaraman B, Finkel T, Stamler JS, Rhee SG and van der
Vliet A. Redox-based regulation of signal transduction:
principles, pitfalls, and promises. Free Radic Biol Med 2008;
45:1-17.
Reddie KG and Carroll KS. Expanding the functional diversity
of proteins through cysteine oxidation. Curr Opin Chem Biol
2008; 12:746-754.
Jacob C, Knight I and Winyard PG. Aspects of the biological
redox chemistry of cysteine: from simple redox responses to
sophisticated signalling pathways. Biol Chem 2006;
387:1385-1397.
Poole LB and Nelson KJ. Discovering mechanisms of
signaling-mediated cysteine oxidation. Curr Opin Chem Biol
2008; 12:18-24.
Ames BN and Shigenaga MK. Oxidants are a major
contributor to aging. Ann N Y Acad Sci 1992; 663:85-96.
Starkov AA. The role of mitochondria in reactive oxygen
species metabolism and signaling. Ann N Y Acad Sci 2008;
1147:37-52.
Lenaz G. Mitochondria and reactive oxygen species. Which
role in physiology and pathology? Adv Exp Med Biol 2012;
942:93-136.
Murphy MP. How mitochondria produce reactive oxygen
species. Biochem J 2009; 417:1-13.
St-Pierre J, Buckingham JA, Roebuck SJ and Brand MD.
Topology of superoxide production from different sites in the
mitochondrial electron transport chain. J Biol Chem 2002;
277:44784-44790.
Miwa S, St-Pierre J, Partridge L and Brand MD. Superoxide
and hydrogen peroxide production by Drosophila
mitochondria. Free Radic Biol Med 2003; 35:938-948.
Drose S and Brandt U. Molecular mechanisms of superoxide
production by the mitochondrial respiratory chain. Adv Exp
Med Biol 2012; 748:145-169.
Manea A. NADPH oxidase-derived reactive oxygen species:
involvement in vascular physiology and pathology. Cell Tissue
Res 2010; 342:325-339.
Bylund J, Brown KL, Movitz C, Dahlgren C and Karlsson A.
Intracellular generation of superoxide by the phagocyte
NADPH oxidase: how, where, and what for? Free Radic Biol
Med 2010; 49:1834-1845.
Harrison R. Physiological roles of xanthine oxidoreductase.
Drug Metab Rev 2004; 36:363-375.
Agarwal A, Banerjee A and Banerjee UC. Xanthine
oxidoreductase: a journey from purine metabolism to
cardiovascular excitation-contraction coupling. Crit Rev
Biotechnol 2011; 31:264-280.
Starkov AA, Fiskum G, Chinopoulos C, Lorenzo BJ, Browne
SE, Patel MS and Beal MF. Mitochondrial alpha-ketoglutarate
dehydrogenase complex generates reactive oxygen species. J
Neurosci 2004; 24:7779-7788.
Ambrus A, Tretter L and Adam-Vizi V. Inhibition of the
alpha-ketoglutarate dehydrogenase-mediated reactive oxygen
species generation by lipoic acid. J Neurochem 2009; 109
Suppl 1:222-229.
Ambrus A, Torocsik B, Tretter L, Ozohanics O and Adam-Vizi
V. Stimulation of reactive oxygen species generation by

ISSN 2168-8761 print/ISSN 2168-877X online

[53]
[54]

[55]

[56]

[57]

[58]

[59]

[60]

[61]

[62]

[63]
[64]
[65]
[66]

[67]

~ 20 ~

disease-causing mutations of lipoamide dehydrogenase. Hum


Mol Genet 2011; 20:2984-2995.
Pollegioni L and Molla G. New biotech applications from
evolved D-amino acid oxidases. Trends Biotechnol 2011;
29:276-283.
Fang J, Sawa T, Akaike T and Maeda H. Tumor-targeted
delivery of polyethylene glycol-conjugated D-amino acid
oxidase for antitumor therapy via enzymatic generation of
hydrogen peroxide. Cancer Res 2002; 62:3138-3143.
Haskew-Layton RE, Payappilly JB, Smirnova NA, Ma TC,
Chan KK, Murphy TH, Guo H, Langley B, Sultana R,
Butterfield DA, Santagata S, Alldred MJ, Gazaryan IG, Bell
GW, Ginsberg SD and Ratan RR. Controlled enzymatic
production of astrocytic hydrogen peroxide protects neurons
from oxidative stress via an Nrf2-independent pathway. Proc
Natl Acad Sci U S A 2010; 107:17385-17390.
Bando Y and Aki K. Mechanisms of generation of oxygen
radicals and reductive mobilization of ferritin iron by
lipoamide dehydrogenase. J Biochem (Tokyo) 1991;
109:450-454.
Sreider CM, Grinblat L and Stoppani AO. Catalysis of
nitrofuran redox-cycling and superoxide anion production by
heart lipoamide dehydrogenase. Biochem Pharmacol 1990;
40:1849-1857.
Gazaryan IG, Krasnikov BF, Ashby GA, Thorneley RN,
Kristal BS and Brown AM. Zinc is a potent inhibitor of thiol
oxidoreductase activity and stimulates reactive oxygen species
production by lipoamide dehydrogenase. J Biol Chem 2002;
277:10064-10072.
Tahara EB, Barros MH, Oliveira GA, Netto LE and
Kowaltowski AJ. Dihydrolipoyl dehydrogenase as a source of
reactive oxygen species inhibited by caloric restriction and
involved in Saccharomyces cerevisiae aging. Faseb J 2007;
21:274-283.
Zhang Q, Zou P, Zhan H, Zhang M, Zhang L, Ge RS and
Huang Y. Dihydrolipoamide dehydrogenase and cAMP are
associated with cadmium-mediated Leydig cell damage.
Toxicol Lett 2011; 205:183-189.
Kareyeva AV, Grivennikova VG, Cecchini G and Vinogradov
AD. Molecular identification of the enzyme responsible for the
mitochondrial
NADH-supported
ammonium-dependent
hydrogen peroxide production. FEBS Lett 2011; 585:385-389.
Kareyeva AV, Grivennikova VG and Vinogradov AD.
Mitochondrial hydrogen peroxide production as determined by
the pyridine nucleotide pool and its redox state. Biochim
Biophys Acta 2012;
Jin RC and Loscalzo J. Vascular Nitric Oxide: Formation and
Function. J Blood Med 2010; 2010:147-162.
Tennyson AG and Lippard SJ. Generation, translocation, and
action of nitric oxide in living systems. Chem Biol 2011;
18:1211-1220.
Feng C. Mechanism of Nitric Oxide Synthase Regulation:
Electron Transfer and Interdomain Interactions. Coord Chem
Rev 2012; 256:393-411.
Hendgen-Cotta UB, Merx MW, Shiva S, Schmitz J, Becher S,
Klare JP, Steinhoff HJ, Goedecke A, Schrader J, Gladwin MT,
Kelm M and Rassaf T. Nitrite reductase activity of myoglobin
regulates respiration and cellular viability in myocardial
ischemia-reperfusion injury. Proc Natl Acad Sci U S A 2008;
105:10256-10261.
Li H, Samouilov A, Liu X and Zweier JL. Characterization of
the magnitude and kinetics of xanthine oxidase-catalyzed
nitrate reduction: evaluation of its role in nitrite and nitric

http://www.researchpub.org/journal/jbpr/jbpr.html

Z. Cai, L.-J. Yan / Journal of Biochemical and Pharmacological Research, Vol. 1 (1): 15-26, March 2013

[68]

[69]

[70]
[71]
[72]
[73]
[74]
[75]
[76]
[77]
[78]
[79]
[80]
[81]

[82]
[83]
[84]
[85]

[86]
[87]

oxide generation in anoxic tissues. Biochemistry 2003;


42:1150-1159.
Sarti P, Forte E, Mastronicola D, Giuffre A and Arese M.
Cytochrome c oxidase and nitric oxide in action: molecular
mechanisms and pathophysiological implications. Biochim
Biophys Acta 2012; 1817:610-619.
Yan LJ, Yang SH, Shu H, Prokai L and Forster MJ.
Histochemical staining and quantification of dihydrolipoamide
dehydrogenase diaphorase activity using blue native PAGE.
Electrophoresis 2007; 28:1036-1045.
Han P, Zhou X, Huang B, Zhang X and Chen C. On-gel
fluorescent visualization and the site identification of
S-nitrosylated proteins. Anal Biochem 2008; 377:150-155.
Huang B and Chen C. Detection of protein S-nitrosation using
irreversible biotinylation procedures (IBP). Free Radic Biol
Med 2010; 49:447-456.
Marsh N and Marsh A. A short history of nitroglycerine and
nitric oxide in pharmacology and physiology. Clin Exp
Pharmacol Physiol 2000; 27:313-319.
Serafim RA, Primi MC, Trossini GH and Ferreira EI. Nitric
oxide: state of the art in drug design. Curr Med Chem 2012;
19:386-405.
Nossaman VE, Nossaman BD and Kadowitz PJ. Nitrates and
nitrites in the treatment of ischemic cardiac disease. Cardiol
Rev 2010; 18:190-197.
Goldstein S and Merenyi G. The chemistry of peroxynitrite:
implications for biological activity. Methods Enzymol 2008;
436:49-61.
Frein D, Schildknecht S, Bachschmid M and Ullrich V. Redox
regulation: a new challenge for pharmacology. Biochem
Pharmacol 2005; 70:811-823.
Poderoso JJ. The formation of peroxynitrite in the applied
physiology of mitochondrial nitric oxide. Arch Biochem
Biophys 2009; 484:214-220.
Castro L, Demicheli V, Tortora V and Radi R. Mitochondrial
protein tyrosine nitration. Free Radic Res 2011; 45:37-52.
Ullrich V and Kissner R. Redox signaling: bioinorganic
chemistry at its best. J Inorg Biochem 2006; 100:2079-2086.
Dalle-Donne I, Aldini G, Carini M, Colombo R, Rossi R and
Milzani A. Protein carbonylation, cellular dysfunction, and
disease progression. J Cell Mol Med 2006; 10:389-406.
Wong PS, Eiserich JP, Reddy S, Lopez CL, Cross CE and van
der Vliet A. Inactivation of glutathione S-transferases by nitric
oxide-derived oxidants: exploring a role for tyrosine nitration.
Arch Biochem Biophys 2001; 394:216-228.
Rao RS and Moller IM. Pattern of occurrence and occupancy
of carbonylation sites in proteins. Proteomics 2011;
11:4166-4173.
Peluffo G and Radi R. Biochemistry of protein tyrosine
nitration in cardiovascular pathology. Cardiovasc Res 2007;
75:291-302.
Szabo C, Ischiropoulos H and Radi R. Peroxynitrite:
biochemistry, pathophysiology and development of
therapeutics. Nat Rev Drug Discov 2007; 6:662-680.
Prokai L, Yan LJ, Vera-Serrano JL, Stevens SM and Forster
MJ. Mass spectrometry-based survey of age-associated protein
carbonylation in rat brain mitochondria. J. Mass Spectrom.
2007; 42:1583-1589.
Madian AG and Regnier FE. Proteomic Identification of
Carbonylated Proteins and Their Oxidation Sites. J Proteome
Res 2010; 9:3766-3780.
Yan LJ and Sohal RS. Prevention of flight activity prolongs the
life span of the housefly, Musca domestica, and attenuates the

ISSN 2168-8761 print/ISSN 2168-877X online

[88]

[89]
[90]
[91]

[92]
[93]
[94]
[95]
[96]

[97]

[98]
[99]
[100]
[101]
[102]
[103]
[104]
[105]

[106]

~ 21 ~

age-associated oxidative damamge to specific mitochondrial


proteins. Free Radic Biol Med 2000; 29:1143-1150.
Yan LJ, Orr WC and Sohal RS. Identification of oxidized
proteins based on sodium dodecyl sulfate-polyacrylamide gel
electrophoresis, immunochemical detection, isoelectric
focusing, and microsequencing. Anal Biochem 1998;
263:67-71.
Yan LJ and Sohal RS. Gel electrophoretic quantitation of
protein carbonyls derivatized with tritiated sodium
borohydride. Anal Biochem 1998; 265:176-182.
Wehr NB and Levine RL. Quantification of protein
carbonylation. Methods Mol Biol 2013; 965:265-281.
Guidi F, Magherini F, Gamberi T, Bini L, Puglia M,
Marzocchini R, Ranaldi F, Modesti PA, Gulisano M and
Modesti A. Plasma protein carbonylation and physical
exercise. Mol Biosyst 2011; 7:640-650.
Wong CM, Cheema AK, Zhang L and Suzuki YJ. Protein
carbonylation as a novel mechanism in redox signaling. Circ
Res 2008; 102:310-318.
Wong CM, Marcocci L, Liu L and Suzuki YJ. Cell signaling
by protein carbonylation and decarbonylation. Antioxid Redox
Signal 2010; 12:393-404.
Curtis JM, Hahn WS, Long EK, Burrill JS, Arriaga EA and
Bernlohr DA. Protein carbonylation and metabolic control
systems. Trends Endocrinol Metab 2012; 23:399-406.
Wong CM, Bansal G, Marcocci L and Suzuki YJ. Proposed
role of primary protein carbonylation in cell signaling. Redox
Rep 2012; 17:90-94.
Serviddio G, Di Venosa N, Federici A, D'Agostino D, Rollo T,
Prigigallo F, Altomare E, Fiore T and Vendemiale G. Brief
hypoxia before normoxic reperfusion (postconditioning)
protects the heart against ischemia-reperfusion injury by
preventing mitochondria peroxyde production and glutathione
depletion. Faseb J 2005; 19:354-361.
Oksala NK, Paimela H, Alhava E and Atalay M. Heat shock
preconditioning induces protein carbonylation and alters
antioxidant protection in superficially injured guinea pig
gastric mucosa in vitro. Dig Dis Sci 2007; 52:1897-1905.
Ferrer-Sueta G and Radi R. Chemical biology of peroxynitrite:
kinetics, diffusion, and radicals. ACS Chem Biol 2009;
4:161-177.
Feeney MB and Schoneich C. Tyrosine modifications in aging.
Antioxid Redox Signal 2012; 17:1571-1579.
Radi R. Protein Tyrosine Nitration: Biochemical Mechanisms
and Structural Basis of Functional Effects. Acc Chem Res
2012;
Robbins RA, Hadeli K, Nelson D, Sato E and Hoyt JC. Nitric
oxide, peroxynitrite, and lower respiratory tract inflammation.
Immunopharmacology 2000; 48:217-221.
Salvemini D, Doyle TM and Cuzzocrea S. Superoxide,
peroxynitrite and oxidative/nitrative stress in inflammation.
Biochem Soc Trans 2006; 34:965-970.
Ohmori H and Kanayama N. Immunogenicity of an
inflammation-associated
product,
tyrosine
nitrated
self-proteins. Autoimmun Rev 2005; 4:224-229.
Sugiura H and Ichinose M. Nitrative stress in inflammatory
lung diseases. Nitric Oxide 2011; 25:138-144.
Cheyuo C, Wu R, Zhou M, Jacob A, Coppa G and Wang P.
Ghrelin suppresses inflammation and neuronal nitric oxide
synthase in focal cerebral ischemia via the vagus nerve. Shock
2011; 35:258-265.
Codoner-Franch P, Tavarez-Alonso S, Murria-Estal R,
Megias-Vericat J, Tortajada-Girbes M and Alonso-Iglesias E.
Nitric oxide production is increased in severely obese children

http://www.researchpub.org/journal/jbpr/jbpr.html

Z. Cai, L.-J. Yan / Journal of Biochemical and Pharmacological Research, Vol. 1 (1): 15-26, March 2013

[107]
[108]
[109]

[110]
[111]
[112]

[113]
[114]
[115]
[116]
[117]
[118]

[119]

[120]

[121]

[122]

[123]
[124]

and related to markers of oxidative stress and inflammation.


Atherosclerosis 2011; 215:475-480.
Rubbo H and Radi R. Protein and lipid nitration: role in redox
signaling and injury. Biochim Biophys Acta 2008;
1780:1318-1324.
Souza JM, Peluffo G and Radi R. Protein tyrosine
nitration--functional alteration or just a biomarker? Free Radic
Biol Med 2008; 45:357-366.
Weber D, Kneschke N, Grimm S, Bergheim I, Breusing N and
Grune T. Rapid and sensitive determination of
protein-nitrotyrosine by ELISA: Application to human plasma.
Free Radic Res 2012; 46:276-285.
Dekker F, Abello N, Wisastra R and Bischoff R. Enrichment
and detection of tyrosine-nitrated proteins. Curr Protoc Protein
Sci 2012; Chapter 14:Unit 14 13.
Webster RP, Roberts VH and Myatt L. Protein nitration in
placenta - functional significance. Placenta 2008; 29:985-994.
Horvath EM, Magenheim R, Kugler E, Vacz G, Szigethy A,
Levardi F, Kollai M, Szabo C and Lacza Z. Nitrative stress and
poly(ADP-ribose) polymerase activation in healthy and
gestational diabetic pregnancies. Diabetologia 2009;
52:1935-1943.
Roos G, Foloppe N and Messens J. Understanding the pK(a) of
Redox Cysteines: The Key Role of Hydrogen Bonding.
Antioxid Redox Signal 2013; 18:94-127.
Chiu J and Dawes IW. Redox control of cell proliferation.
Trends Cell Biol 2012; 22:592-601.
Broniowska KA and Hogg N. The chemical biology of
S-nitrosothiols. Antioxid Redox Signal 2012; 17:969-980.
Rhee SG, Chang TS, Bae YS, Lee SR and Kang SW. Cellular
regulation by hydrogen peroxide. J Am Soc Nephrol 2003;
14:S211-S215.
Rhee SG, Bae YS, Lee SR and Kwon J. Hydrogen peroxide: a
key messenger that modulates protein phosphorylation through
cysteine oxidation. Sci STKE 2000; 2000:pe1.
Lim JC, Gruschus JM, Kim G, Berlett BS, Tjandra N and
Levine RL. A low pKa cysteine at the active site of mouse
methionine sulfoxide reductase A. J Biol Chem 2012;
287:25596-25601.
Antoine M, Gand A, Boschi-Muller S and Branlant G.
Characterization of the amino acids from Neisseria
meningitidis MsrA involved in the chemical catalysis of the
methionine sulfoxide reduction step. J Biol Chem 2006;
281:39062-39070.
Sengupta S, Chen H, Togawa T, DiBello PM, Majors AK,
Budy B, Ketterer ME and Jacobsen DW. Albumin thiolate
anion is an intermediate in the formation of
albumin-S-S-homocysteine.
J
Biol
Chem
2001;
276:30111-30117.
Turell L, Botti H, Carballal S, Ferrer-Sueta G, Souza JM,
Duran R, Freeman BA, Radi R and Alvarez B. Reactivity of
sulfenic acid in human serum albumin. Biochemistry 2008;
47:358-367.
Turell L, Botti H, Carballal S, Radi R and Alvarez B. Sulfenic
acid--a key intermediate in albumin thiol oxidation. J
Chromatogr B Analyt Technol Biomed Life Sci 2009;
877:3384-3392.
Alvarez B, Carballal S, Turell L and Radi R. Formation and
reactions of sulfenic acid in human serum albumin. Methods
Enzymol 2010; 473:117-136.
Winterbourn CC and Hampton MB. Thiol chemistry and
specificity in redox signaling. Free Radic Biol Med 2008;
45:549-561.

ISSN 2168-8761 print/ISSN 2168-877X online

[125] Riederer BM. Oxidation proteomics: The role of thiol


modifications. Current Proteomics 2009; 6:51-62.
[126] Murphy MP. Mitochondrial thiols in antioxidant protection
and redox signaling: distinct roles for glutathionylation and
other thiol modifications. Antioxid Redox Signal 2012;
16:476-495.
[127] Go YM, Halvey PJ, Hansen JM, Reed M, Pohl J and Jones DP.
Reactive aldehyde modification of thioredoxin-1 activates
early steps of inflammation and cell adhesion. Am J Pathol
2007; 171:1670-1681.
[128] Carvalho CM, Chew EH, Hashemy SI, Lu J and Holmgren A.
Inhibition of the human thioredoxin system. A molecular
mechanism of mercury toxicity. J Biol Chem 2008;
283:11913-11923.
[129] Salsbury FR, Jr., Knutson ST, Poole LB and Fetrow JS.
Functional site profiling and electrostatic analysis of cysteines
modifiable to cysteine sulfenic acid. Protein Sci 2008;
17:299-312.
[130] Jacob C, Battaglia E, Burkholz T, Peng D, Bagrel D and
Montenarh M. Control of oxidative posttranslational cysteine
modifications: from intricate chemistry to widespread
biological and medical applications. Chem Res Toxicol 2012;
25:588-604.
[131] Biswas S, Chida AS and Rahman I. Redox modifications of
protein-thiols: emerging roles in cell signaling. Biochem
Pharmacol 2006; 71:551-564.
[132] Requejo R, Hurd TR, Costa NJ and Murphy MP. Cysteine
residues exposed on protein surfaces are the dominant
intramitochondrial thiol and may protect against oxidative
damage. Febs J 2010; 277:1465-1480.
[133] Chen YY, Chu HM, Pan KT, Teng CH, Wang DL, Wang AH,
Khoo KH and Meng TC. Cysteine S-nitrosylation protects
protein-tyrosine phosphatase 1B against oxidation-induced
permanent inactivation. J Biol Chem 2008; 283:35265-35272.
[134] Lou YW, Chen YY, Hsu SF, Chen RK, Lee CL, Khoo KH,
Tonks NK and Meng TC. Redox regulation of the protein
tyrosine phosphatase PTP1B in cancer cells. Febs J 2008;
275:69-88.
[135] Madian AG, Hindupur J, Hulleman JD, Diaz-Maldonado N,
Mishra VR, Guigard E, Kay CM, Rochet JC and Regnier FE.
Effect of single amino acid substitution on oxidative
modifications of the Parkinson's disease-related protein, DJ-1.
Mol Cell Proteomics 2012; 11:M111 010892.
[136] Wang Y, Yang J and Yi J. Redox sensing by proteins:
oxidative modifications on cysteines and the consequent
events. Antioxid Redox Signal 2012; 16:649-657.
[137] Poole LB, Karplus PA and Claiborne A. Protein sulfenic acids
in redox signaling. Annu Rev Pharmacol Toxicol 2004;
44:325-347.
[138] van Montfort RL, Congreve M, Tisi D, Carr R and Jhoti H.
Oxidation state of the active-site cysteine in protein tyrosine
phosphatase 1B. Nature 2003; 423:773-777.
[139] Mansuy D and Dansette PM. Sulfenic acids as reactive
intermediates in xenobiotic metabolism. Arch Biochem
Biophys 2011; 507:174-185.
[140] Roos G and Messens J. Protein sulfenic acid formation: from
cellular damage to redox regulation. Free Radic Biol Med
2011; 51:314-326.
[141] Kettenhofen NJ and Wood MJ. Formation, reactivity, and
detection of protein sulfenic acids. Chem Res Toxicol 2010;
23:1633-1646.
[142] Saurin AT, Neubert H, Brennan JP and Eaton P. Widespread
sulfenic acid formation in tissues in response to hydrogen
peroxide. Proc Natl Acad Sci U S A 2004; 101:17982-17987.

~ 22 ~

http://www.researchpub.org/journal/jbpr/jbpr.html

Z. Cai, L.-J. Yan / Journal of Biochemical and Pharmacological Research, Vol. 1 (1): 15-26, March 2013

[143] Charles RL, Schroder E, May G, Free P, Gaffney PR, Wait R,


Begum S, Heads RJ and Eaton P. Protein sulfenation as a redox
sensor: proteomics studies using a novel biotinylated
dimedone analogue. Mol Cell Proteomics 2007; 6:1473-1484.
[144] Yan LJ, Sumien N, Thangthaeng N and Forster MJ. Reversible
inactivation of dihydrolipoamide dehydrogenase by
mitochondrial hydrogen peroxide. Free Radic Res 2013;
47:123-133.
[145] Poole LB. Formation and functions of protein sulfenic acids.
Curr Protoc Toxicol 2004; Chapter 17:Unit17 1.
[146] Depuydt M, Leonard SE, Vertommen D, Denoncin K,
Morsomme P, Wahni K, Messens J, Carroll KS and Collet JF.
A periplasmic reducing system protects single cysteine
residues from oxidation. Science 2009; 326:1109-1111.
[147] Takanishi CL, Ma LH and Wood MJ. A genetically encoded
probe for cysteine sulfenic acid protein modification in vivo.
Biochemistry 2007; 46:14725-14732.
[148] Poole LB, Klomsiri C, Knaggs SA, Furdui CM, Nelson KJ,
Thomas MJ, Fetrow JS, Daniel LW and King SB. Fluorescent
and affinity-based tools to detect cysteine sulfenic acid
formation in proteins. Bioconjug Chem 2007; 18:2004-2017.
[149] Seo YH and Carroll KS. Quantification of protein sulfenic acid
modifications
using
isotope-coded
dimedone
and
iododimedone. Angew Chem Int Ed Engl 2011; 50:1342-1345.
[150] Seo YH and Carroll KS. Profiling protein thiol oxidation in
tumor cells using sulfenic acid-specific antibodies. Proc Natl
Acad Sci U S A 2009; 106:16163-16168.
[151] Kaiserova K, Srivastava S, Hoetker JD, Awe SO, Tang XL,
Cai J and Bhatnagar A. Redox activation of aldose reductase in
the ischemic heart. J Biol Chem 2006; 281:15110-15120.
[152] Kaiserova K, Tang XL, Srivastava S and Bhatnagar A. Role of
nitric oxide in regulating aldose reductase activation in the
ischemic heart. J Biol Chem 2008; 283:9101-9112.
[153] Wetzelberger K, Baba SP, Thirunavukkarasu M, Ho YS,
Maulik N, Barski OA, Conklin DJ and Bhatnagar A.
Postischemic deactivation of cardiac aldose reductase: role of
glutathione S-transferase P and glutaredoxin in regeneration of
reduced thiols from sulfenic acids. J Biol Chem 2010;
285:26135-26148.
[154] Michalek RD, Nelson KJ, Holbrook BC, Yi JS, Stridiron D,
Daniel LW, Fetrow JS, King SB, Poole LB and Grayson JM.
The requirement of reversible cysteine sulfenic acid formation
for T cell activation and function. J Immunol 2007;
179:6456-6467.
[155] Murakami T, Nojiri M, Nakayama H, Odaka M, Yohda M,
Dohmae N, Takio K, Nagamune T and Endo I.
Post-translational modification is essential for catalytic activity
of nitrile hydratase. Protein Sci 2000; 9:1024-1030.
[156] Foster MW, McMahon TJ and Stamler JS. S-nitrosylation in
health and disease. Trends Mol Med 2003; 9:160-168.
[157] Foster MW, Hess DT and Stamler JS. Protein S-nitrosylation
in health and disease: a current perspective. Trends Mol Med
2009; 15:391-404.
[158] Wang Y, Yun BW, Kwon E, Hong JK, Yoon J and Loake GJ.
S-nitrosylation: an emerging redox-based post-translational
modification in plants. J Exp Bot 2006; 57:1777-1784.
[159] Akhtar MW, Sunico CR, Nakamura T and Lipton SA. Redox
Regulation of Protein Function via Cysteine S-Nitrosylation
and Its Relevance to Neurodegenerative Diseases. Int J Cell
Biol 2012; 2012:463756.
[160] Sun J and Murphy E. Protein S-nitrosylation and
cardioprotection. Circ Res 2010; 106:285-296.

ISSN 2168-8761 print/ISSN 2168-877X online

[161] Okado-Matsumoto A and Fridovich I. Putative denitrosylase


activity of Cu,Zn-superoxide dismutase. Free Radic Biol Med
2007; 43:830-836.
[162] Duan S and Chen C. S-nitrosylation/denitrosylation and
apoptosis of immune cells. Cell Mol Immunol 2007;
4:353-358.
[163] Foster MW, Yang Z, Gooden DM, Thompson JW, Ball CH,
Turner ME, Hou Y, Pi J, Moseley MA and Que LG. Proteomic
characterization of the cellular response to nitrosative stress
mediated by s-nitrosoglutathione reductase inhibition. J
Proteome Res 2012; 11:2480-2491.
[164] Beigi F, Gonzalez DR, Minhas KM, Sun QA, Foster MW,
Khan SA, Treuer AV, Dulce RA, Harrison RW, Saraiva RM,
Premer C, Schulman IH, Stamler JS and Hare JM. Dynamic
denitrosylation via S-nitrosoglutathione reductase regulates
cardiovascular function. Proc Natl Acad Sci U S A 2012;
109:4314-4319.
[165] Barone E, Trombino S, Cassano R, Sgambato A, De Paola B,
Di Stasio E, Picci N, Preziosi P and Mancuso C.
Characterization of the S-denitrosylating activity of bilirubin. J
Cell Mol Med 2009; 13:2365-2375.
[166] Yan LJ, Liu L and Forster MJ. Reversible inactivation of
dihydrolipoamide dehydrogenase by Angeli's salt. Acta
Biophysica Sinica (Sheng Wu Wu Li Hsueh Bao) 2012;
28:341-350.
[167] Haldar SM and Stamler JS. S-nitrosylation: integrator of
cardiovascular performance and oxygen delivery. J Clin Invest
2013; 123:101-110.
[168] Sun J, Morgan M, Shen RF, Steenbergen C and Murphy E.
Preconditioning results in S-nitrosylation of proteins involved
in regulation of mitochondrial energetics and calcium
transport. Circ Res 2007; 101:1155-1163.
[169] Lin J, Steenbergen C, Murphy E and Sun J. Estrogen
receptor-beta activation results in S-nitrosylation of proteins
involved in cardioprotection. Circulation 2009; 120:245-254.
[170] Murphy E, Kohr M, Sun J, Nguyen T and Steenbergen C.
S-nitrosylation: a radical way to protect the heart. J Mol Cell
Cardiol 2012; 52:568-577.
[171] Sheng H, Reynolds JD, Auten RL, Demchenko IT, Piantadosi
CA, Stamler JS and Warner DS. Pharmacologically augmented
S-nitrosylated hemoglobin improves recovery from murine
subarachnoid hemorrhage. Stroke 2011; 42:471-476.
[172] Penna C, Perrelli MG, Tullio F, Moro F, Parisella ML, Merlino
A and Pagliaro P. Post-ischemic early acidosis in cardiac
postconditioning modifies the activity of antioxidant enzymes,
reduces nitration, and favors protein S-nitrosylation. Pflugers
Arch 2011; 462:219-233.
[173] Xiong Y, Uys JD, Tew KD and Townsend DM.
S-glutathionylation: from molecular mechanisms to health
outcomes. Antioxid Redox Signal 2011; 15:233-270.
[174] Pastore A and Piemonte F. S-Glutathionylation signaling in
cell biology: progress and prospects. Eur J Pharm Sci 2012;
46:279-292.
[175] Pimentel D, Haeussler DJ, Matsui R, Burgoyne JR, Cohen RA
and Bachschmid MM. Regulation of cell physiology and
pathology by protein S-glutathionylation: lessons learned from
the cardiovascular system. Antioxid Redox Signal 2012;
16:524-542.
[176] Zweier JL, Chen CA and Druhan LJ. S-glutathionylation
reshapes our understanding of endothelial nitric oxide synthase
uncoupling and nitric oxide/reactive oxygen species-mediated
signaling. Antioxid Redox Signal 2011; 14:1769-1775.

~ 23 ~

http://www.researchpub.org/journal/jbpr/jbpr.html

Z. Cai, L.-J. Yan / Journal of Biochemical and Pharmacological Research, Vol. 1 (1): 15-26, March 2013

[177] Dalle-Donne I, Rossi R, Colombo G, Giustarini D and Milzani


A. Protein S-glutathionylation: a regulatory device from
bacteria to humans. Trends Biochem Sci 2009; 34:85-96.
[178] Dalle-Donne I, Colombo G, Gagliano N, Colombo R,
Giustarini D, Rossi R and Milzani A. S-glutathiolation in life
and death decisions of the cell. Free Radic Res 2011; 45:3-15.
[179] Velu CS, Niture SK, Doneanu CE, Pattabiraman N and
Srivenugopal KS. Human p53 is inhibited by glutathionylation
of cysteines present in the proximal DNA-binding domain
during oxidative stress. Biochemistry 2007; 46:7765-7780.
[180] Xie Y, Kole S, Precht P, Pazin MJ and Bernier M.
S-glutathionylation impairs signal transducer and activator of
transcription 3 activation and signaling. Endocrinology 2009;
150:1122-1131.
[181] Kambe T, Song T, Takata T, Hatano N, Miyamoto Y, Nozaki
N, Naito Y, Tokumitsu H and Watanabe Y. Inactivation of
Ca2+/calmodulin-dependent
protein
kinase
I
by
S-glutathionylation of the active-site cysteine residue. FEBS
Lett 2010; 584:2478-2484.
[182] Kim YJ, Kim D, Illuzzi JL, Delaplane S, Su D, Bernier M,
Gross ML, Georgiadis MM and Wilson DM, 3rd.
S-glutathionylation of cysteine 99 in the APE1 protein impairs
abasic endonuclease activity. J Mol Biol 2011; 414:313-326.
[183] Coan C, Ji JY, Hideg K and Mehlhorn RJ. Protein sulfhydryls
are protected from irreversible oxidation by conversion to
mixed disulfides. Arch Biochem Biophys 1992; 295:369-378.
[184] Chae HZ, Oubrahim H, Park JW, Rhee SG and Chock PB.
Protein glutathionylation in the regulation of peroxiredoxins: a
family of thiol-specific peroxidases that function as
antioxidants, molecular chaperones, and signal modulators.
Antioxid Redox Signal 2012; 16:506-523.
[185] Cooper AJ, Pinto JT and Callery PS. Reversible and
irreversible protein glutathionylation: biological and clinical
aspects. Expert Opin Drug Metab Toxicol 2011; 7:891-910.
[186] McLain AL, Szweda PA and Szweda LI. alpha-Ketoglutarate
dehydrogenase: a mitochondrial redox sensor. Free Radic Res
2011; 45:29-36.
[187] Rasmussen HH, Hamilton EJ, Liu CC and Figtree GA.
Reversible oxidative modification: implications for
cardiovascular physiology and pathophysiology. Trends
Cardiovasc Med 2010; 20:85-90.
[188] Sakai J, Li J, Subramanian KK, Mondal S, Bajrami B, Hattori
H, Jia Y, Dickinson BC, Zhong J, Ye K, Chang CJ, Ho YS,
Zhou J and Luo HR. Reactive oxygen species-induced actin
glutathionylation controls actin dynamics in neutrophils.
Immunity 2012; 37:1037-1049.
[189] Pfefferie A, Mailloux RJ, Adjeitey CN and Harper ME.
Glutathionylation of UCP2 sensitizes drug resistant leukemia
cells to chemotherapeutics. Biochim Biophys Acta 2013;
1833:80-89.
[190] Queiroga CS, Almeida AS, Martel C, Brenner C, Alves PM
and Vieira HL. Glutathionylation of adenine nucleotide
translocase induced by carbon monoxide prevents
mitochondrial membrane permeabilization and apoptosis. J
Biol Chem 2010; 285:17077-17088.
[191] O'Brian CA and Chu F. Post-translational disulfide
modifications in cell signaling--role of inter-protein,
intra-protein, S-glutathionyl, and S-cysteaminyl disulfide
modifications in signal transmission. Free Radic Res 2005;
39:471-480.
[192] Huang KP and Huang FL. Glutathionylation of proteins by
glutathione disulfide S-oxide. Biochem Pharmacol 2002;
64:1049-1056.

ISSN 2168-8761 print/ISSN 2168-877X online

[193] Cotgreave IA. Analytical developments in the assay of intraand extracellular GSH homeostasis: specific protein
S-glutathionylation, cellular GSH and mixed disulphide
compartmentalisation and interstitial GSH redox balance.
Biofactors 2003; 17:269-277.
[194] Xiao R, Wilkinson B, Solovyov A, Winther JR, Holmgren A,
Lundstrom-Ljung J and Gilbert HF. The contributions of
protein disulfide isomerase and its homologues to oxidative
protein folding in the yeast endoplasmic reticulum. J Biol
Chem 2004; 279:49780-49786.
[195] Narayan M. Disulfide bonds: protein folding and subcellular
protein trafficking. Febs J 2012; 279:2272-2282.
[196] Benham AM. The protein disulfide isomerase family: key
players in health and disease. Antioxid Redox Signal 2012;
16:781-789.
[197] Depuydt M, Messens J and Collet JF. How proteins form
disulfide bonds. Antioxid Redox Signal 2011; 15:49-66.
[198] Sommer A and Traut RR. Diagonal polyacrylamide-dodecyl
sulfate gel electrophoresis for the identification of ribosomal
proteins crosslinked with methyl-4-mercaptobutyrimidate.
Proc Natl Acad Sci U S A 1974; 71:3946-3950.
[199] Cumming RC, Andon NL, Haynes PA, Park M, Fischer WH
and Schubert D. Protein disulfide bond formation in the
cytoplasm during oxidative stress. J Biol Chem 2004;
279:21749-21758.
[200] Danciu TE and Whitman M. Oxidative stress drives disulfide
bond formation between basic helix-loop-helix transcription
factors. J Cell Biochem 2010; 109:417-424.
[201] McDonagh B and Sheehan D. Effect of oxidative stress on
protein thiols in the blue mussel Mytilus edulis: proteomic
identification of target proteins. Proteomics 2007;
7:3395-3403.
[202] McDonagh B and Sheehan D. Effects of oxidative stress on
protein thiols and disulphides in Mytilus edulis revealed by
proteomics: actin and protein disulphide isomerase are redox
targets. Mar Environ Res 2008; 66:193-195.
[203] An BC, Lee SS, Lee EM, Wi SG, Park W and Chung BY.
Global analysis of disulfide bond proteins in Pseudomonas
aeruginosa exposed to hydrogen peroxide and gamma rays. Int
J Radiat Biol 2010; 86:400-408.
[204] Leichert LI, Gehrke F, Gudiseva HV, Blackwell T, Ilbert M,
Walker AK, Strahler JR, Andrews PC and Jakob U.
Quantifying changes in the thiol redox proteome upon
oxidative stress in vivo. Proc Natl Acad Sci U S A 2008;
105:8197-8202.
[205] McDonagh B. Diagonal electrophoresis for the detection of
protein disulfides. Methods Mol Biol 2012; 309-315.
[206] Ghezzi P, Bonetto V and Fratelli M. Thiol-disulfide balance:
from the concept of oxidative stress to that of redox regulation.
Antioxid Redox Signal 2005; 7:964-972.
[207] Shimizu Y and Hendershot LM. Oxidative folding: cellular
strategies for dealing with the resultant equimolar production
of reactive oxygen species. Antioxid Redox Signal 2009;
11:2317-2331.
[208] Guo Z, Kozlov S, Lavin MF, Person MD and Paull TT. ATM
activation by oxidative stress. Science 2010; 330:517-521.
[209] Li W, Zhang J and An W. The conserved CXXC motif of
hepatic stimulator substance is essential for its role in
mitochondrial protection in H2O2-induced cell apoptosis.
FEBS Lett 2010; 584:3929-3935.
[210] Wei PC, Hsieh YH, Su MI, Jiang XJ, Hsu PH, Lo WT, Weng
JY, Jeng YM, Wang JM, Chen PL, Chang YC, Lee KF, Tsai
MD, Shew JY and Lee WH. Loss of the Oxidative Stress

~ 24 ~

http://www.researchpub.org/journal/jbpr/jbpr.html

Z. Cai, L.-J. Yan / Journal of Biochemical and Pharmacological Research, Vol. 1 (1): 15-26, March 2013

[211]
[212]

[213]

[214]

[215]
[216]
[217]

[218]
[219]

[220]
[221]

[222]

[223]

[224]

[225]
[226]

Sensor NPGPx Compromises GRP78 Chaperone Activity and


Induces Systemic Disease. Mol Cell 2012;
Savitsky PA and Finkel T. Redox regulation of Cdc25C. J Biol
Chem 2002; 277:20535-20540.
Li XH, Chen Z, Gao YS, Yan YB, Zhang F, Meng FG and
Zhou HM. Generation of the oxidized form protects human
brain type creatine kinase against cystine-induced inactivation.
Int J Biol Macromol 2011; 48:239-242.
Dinkova-Kostova AT, Holtzclaw WD, Cole RN, Itoh K,
Wakabayashi N, Katoh Y, Yamamoto M and Talalay P. Direct
evidence that sulfhydryl groups of Keap1 are the sensors
regulating induction of phase 2 enzymes that protect against
carcinogens and oxidants. Proc Natl Acad Sci U S A 2002;
99:11908-11913.
Wakabayashi N, Dinkova-Kostova AT, Holtzclaw WD, Kang
MI, Kobayashi A, Yamamoto M, Kensler TW and Talalay P.
Protection against electrophile and oxidant stress by induction
of the phase 2 response: fate of cysteines of the Keap1 sensor
modified by inducers. Proc Natl Acad Sci U S A 2004;
101:2040-2045.
Kensler TW, Wakabayashi N and Biswal S. Cell survival
responses to environmental stresses via the Keap1-Nrf2-ARE
pathway. Annu Rev Pharmacol Toxicol 2007; 47:89-116.
Dinkova-Kostova AT and Talalay P. Direct and indirect
antioxidant properties of inducers of cytoprotective proteins.
Mol Nutr Food Res 2008; 52 Suppl 1:S128-S138.
Villeneuve NF, Lau A and Zhang DD. Regulation of the
Nrf2-Keap1 antioxidant response by the ubiquitin proteasome
system: an insight into cullin-ring ubiquitin ligases. Antioxid
Redox Signal 2011; 13:1699-1712.
Yan W, Wang HD, Hu ZG, Wang QF and Yin HX. Activation
of Nrf2-ARE pathway in brain after traumatic brain injury.
Neurosci Lett 2008; 431:150-154.
Abbas K, Breton J, Planson AG, Bouton C, Bignon J, Seguin
C, Riquier S, Toledano MB and Drapier JC. Nitric oxide
activates an Nrf2/sulfiredoxin antioxidant pathway in
macrophages. Free Radic Biol Med 2011; 51:107-114.
Baird L and Dinkova-Kostova AT. The cytoprotective role of
the Keap1-Nrf2 pathway. Arch Toxicol 2011; 85:241-272.
Bell KF, Al-Mubarak B, Fowler JH, Baxter PS, Gupta K,
Tsujita T, Chowdhry S, Patani R, Chandran S, Horsburgh K,
Hayes JD and Hardingham GE. Mild oxidative stress activates
Nrf2 in astrocytes, which contributes to neuroprotective
ischemic preconditioning. Proc Natl Acad Sci U S A 2011;
108:E1-2; author reply E3-4.
Bell KFS, Fowler JH, Al-Mubarak B, Horsburgh K and
Hardingham GE. Activation of Nrf2-Regulated Glutathione
Pathway
Genes
by
Ischemic
Preconditioning
10.1155/2011/689524. Oxidative Medicine and Cellular
Longevity 2011; 2011:
Li MH, Cha YN and Surh YJ. Peroxynitrite induces HO-1
expression
via
PI3K/Akt-dependent
activation
of
NF-E2-related factor 2 in PC12 cells. Free Radic Biol Med
2006; 41:1079-1091.
Zheng H, Whitman SA, Wu W, Wondrak GT, Wong PK, Fang
D and Zhang DD. Therapeutic potential of Nrf2 activators in
streptozotocin-induced diabetic nephropathy. Diabetes 2011;
60:3055-3066.
Eaton P, Bell RM, Cave AC and Shattock MJ. Ischemic
preconditioning: a potential role for protein S-thiolation?
Antioxid Redox Signal 2005; 7:882-888.
Blanco M, Lizasoain I, Sobrino T, Vivancos J and Castillo J.
Ischemic preconditioning: A novel target for neuroprotective
therapy. Cerebrovasc Dis 2006; 21 Suppl 2:38-47.

ISSN 2168-8761 print/ISSN 2168-877X online

[227] Sun J, Steenbergen C and Murphy E. S-nitrosylation:


NO-related redox signaling to protect against oxidative stress.
Antioxid Redox Signal 2006; 8:1693-1705.
[228] Sun J. Protein s-nitrosylation: A role of nitric oxide signaling
in cardiac ischemic preconditioning. Acta Physiologica Sinica
2007; 59:544-552.
[229] Shiva S, Sack MN, Greer JJ, Duranski M, Ringwood LA,
Burwell L, Wang X, MacArthur PH, Shoja A, Raghavachari N,
Calvert JW, Brookes PS, Lefer DJ and Gladwin MT. Nitrite
augments tolerance to ischemia/reperfusion injury via the
modulation of mitochondrial electron transfer. J Exp Med
2007; 204:2089-2102.
[230] Kohr MJ, Sun J, Aponte A, Wang G, Gucek M, Murphy E and
Steenbergen C. Simultaneous measurement of protein
oxidation and S-nitrosylation during preconditioning and
ischemia/reperfusion injury with resin-assisted capture. Circ
Res 2011; 108:418-426.
[231] Hill BG and Darley-Usmar VM. S-nitrosation and thiol
switching in the mitochondrion: a new paradigm for
cardioprotection in ischaemic preconditioning. Biochem J
2008; 412:e11-e13.
[232] Kitagawa K, Matsumoto M, Tagaya M, Hata R, Ueda H,
Niinobe M, Handa N, Fukunaga R, Kimura K, Mikoshiba K
and et al. 'Ischemic tolerance' phenomenon found in the brain.
Brain Res 1990; 528:21-24.
[233] Yang CW, Ahn HJ, Han HJ, Kim WY, Li C, Shin MJ, Kim SK,
Park JH, Kim YS, Moon IS and Bang BK. Pharmacological
preconditioning with low-dose cyclosporine or FK506 reduces
subsequent ischemia/reperfusion injury in rat kidney.
Transplantation 2001; 72:1753-1759.
[234] Ravingerova T. Intrinsic defensive mechanisms in the heart: a
potential novel approach to cardiac protection against ischemic
injury. Gen Physiol Biophys 2007; 26:3-13.
[235] Stetler RA, Zhang F, Liu C and Chen J. Ischemic tolerance as
an active and intrinsic neuroprotective mechanism. Handb Clin
Neurol 2009; 92:171-195.
[236] Kitagawa K. Ischemic tolerance in the brain: endogenous
adaptive machinery against ischemic stress. J Neurosci Res
2012; 90:1043-1054.
[237] Della-Morte D, Guadagni F, Palmirotta R, Ferroni P, Testa G,
Cacciatore F, Abete P, Rengo F, Perez-Pinzon MA, Sacco RL
and Rundek T. Genetics and genomics of ischemic tolerance:
focus on cardiac and cerebral ischemic preconditioning.
Pharmacogenomics 2012; 13:1741-1757.
[238] Yang X, Cohen MV and Downey JM. Mechanism of
cardioprotection by early ischemic preconditioning.
Cardiovasc Drugs Ther 2010; 24:225-234.
[239] Vanden Hoek TL, Becker LB, Shao Z, Li C and Schumacker
PT. Reactive oxygen species released from mitochondria
during
brief
hypoxia
induce
preconditioning in
cardiomyocytes. J Biol Chem 1998; 273:18092-18098.
[240] da Silva MM, Sartori A, Belisle E and Kowaltowski AJ.
Ischemic preconditioning inhibits mitochondrial respiration,
increases H2O2 release, and enhances K+ transport. Am J
Physiol Heart Circ Physiol 2003; 285:H154-H162.
[241] Kunz A, Park L, Abe T, Gallo EF, Anrather J, Zhou P and
Iadecola C. Neurovascular protection by ischemic tolerance:
role of nitric oxide and reactive oxygen species. J Neurosci
2007; 27:7083-7093.
[242] Liu Y, Yang XM, Iliodromitis EK, Kremastinos DT, Dost T,
Cohen MV and Downey JM. Redox signaling at reperfusion is
required for protection from ischemic preconditioning but not
from a direct PKC activator. Basic Res Cardiol 2008;
103:54-59.

~ 25 ~

http://www.researchpub.org/journal/jbpr/jbpr.html

Z. Cai, L.-J. Yan / Journal of Biochemical and Pharmacological Research, Vol. 1 (1): 15-26, March 2013

[243] Penna C, Mancardi D, Rastaldo R and Pagliaro P.


Cardioprotection: a radical view Free radicals in pre and
postconditioning. Biochim Biophys Acta 2009; 1787:781-793.
[244] Geracitano R, Tozzi A, Berretta N, Florenzano F, Guatteo E,
Viscomi MT, Chiolo B, Molinari M, Bernardi G and Mercuri
NB. Protective role of hydrogen peroxide in oxygen-deprived
dopaminergic neurones of the rat substantia nigra. J Physiol
2005; 568:97-110.
[245] Nistico R, Piccirilli S, Cucchiaroni ML, Armogida M, Guatteo
E, Giampa C, Fusco FR, Bernardi G, Nistico G and Mercuri
NB. Neuroprotective effect of hydrogen peroxide on an in vitro
model of brain ischaemia. Br J Pharmacol 2008;
153:1022-1029.
[246] Chadwick W, Zhou Y, Park SS, Wang L, Mitchell N, Stone
MD, Becker KG, Martin B and Maudsley S. Minimal peroxide
exposure of neuronal cells induces multifaceted adaptive
responses. PLoS One 2010; 5:e14352.
[247] Zhao H, Sapolsky RM and Steinberg GK. Interrupting
reperfusion as a stroke therapy: ischemic postconditioning
reduces infarct size after focal ischemia in rats. J Cereb Blood
Flow Metab 2006; 26:1114-1121.
[248] Zhao H. The protective effect of ischemic postconditioning
against ischemic injury: from the heart to the brain. J
Neuroimmune Pharmacol 2007; 2:313-318.
[249] Zhao H. Ischemic postconditioning as a novel avenue to
protect against brain injury after stroke. J Cereb Blood Flow
Metab 2009; 29:873-885.
[250] Zhao H, Ren C, Chen X and Shen J. From rapid to delayed and
remote postconditioning: the evolving concept of ischemic
postconditioning in brain ischemia. Curr Drug Targets 2012;
13:173-187.
[251] Hausenloy DJ and Yellon DM. Survival kinases in ischemic
preconditioning and postconditioning. Cardiovasc Res 2006;
70:240-253.
[252] Lim SY, Davidson SM, Hausenloy DJ and Yellon DM.
Preconditioning and postconditioning: the essential role of the
mitochondrial permeability transition pore. Cardiovasc Res
2007; 75:530-535.
[253] Hausenloy DJ, Tsang A and Yellon DM. The reperfusion
injury salvage kinase pathway: a common target for both
ischemic preconditioning and postconditioning. Trends
Cardiovasc Med 2005; 15:69-75.
[254] Pignataro G, Scorziello A, Di Renzo G and Annunziato L.
Post-ischemic brain damage: effect
of
ischemic

ISSN 2168-8761 print/ISSN 2168-877X online

[255]
[256]

[257]
[258]
[259]
[260]
[261]

[262]
[263]
[264]
[265]

[266]
[267]

~ 26 ~

preconditioning and postconditioning and identification of


potential candidates for stroke therapy. Febs J 2009;
276:46-57.
Zhao ZQ and Vinten-Johansen J. Postconditioning: reduction
of reperfusion-induced injury. Cardiovasc Res 2006;
70:200-211.
Rodriguez-Sinovas A, Abdallah Y, Piper HM and
Garcia-Dorado D. Reperfusion injury as a therapeutic
challenge in patients with acute myocardial infarction. Heart
Fail Rev 2007; 12:207-216.
Gursoy-Ozdemir Y, Yemisci M and Dalkara T. Microvascular
protection is essential for successful neuroprotection in stroke.
J Neurochem 2012; 123 Suppl 2:2-11.
Hausenloy DJ and Yellon DM. Preconditioning and
postconditioning: united at reperfusion. Pharmacol Ther 2007;
116:173-191.
Hausenloy DJ, Wynne AM and Yellon DM. Ischemic
preconditioning targets the reperfusion phase. Basic Res
Cardiol 2007; 102:445-452.
Burwell LS, Nadtochiy SM and Brookes PS. Cardioprotection
by metabolic shut-down and gradual wake-up. J Mol Cell
Cardiol 2009; 46:804-810.
Muller BA and Dhalla NS. Mechanisms of the beneficial
actions of ischemic preconditioning on subcellular remodeling
in ischemic-reperfused heart. Curr Cardiol Rev 2010;
6:255-264.
Murillo D, Kamga C, Mo L and Shiva S. Nitrite as a mediator
of ischemic preconditioning and cytoprotection. Nitric Oxide
2011; 25:70-80.
Wall SB, Oh JY, Diers AR and Landar A. Oxidative
modification of proteins: an emerging mechanism of cell
signaling. Front Physiol 2012; 3:369.
Finkel T. Signal transduction by reactive oxygen species. J
Cell Biol 2011; 194:7-15.
Chung HS, Wang SB, Venkatraman V, Murray CI and Van
Eyk JE. Cysteine oxidative posttranslational modifications:
emerging regulation in the cardiovascular system. Circ Res
2013; 112:382-392.
Ma Q. Role of nrf2 in oxidative stress and toxicity. Annu Rev
Pharmacol Toxicol 2013; 53:401-426.
Pagliaro P, Moro F, Tullio F, Perrelli MG and Penna C.
Cardioprotective pathways during reperfusion: focus on redox
signaling and other modalities of cell signaling. Antioxid
Redox Signal 2011; 14:833-850.

http://www.researchpub.org/journal/jbpr/jbpr.html

Anda mungkin juga menyukai